c 2006 Tech Science Press Copyright CMES, vol.14, no.2, pp.91-100, 2006 A lattice-based cell model for calculating thermal capacity and expansion of single wall carbon nanotubes Xianwu Ling1 and S.N. Atluri Abstract: In this paper, a lattice-based cell model is proposed for single wall carbon nanotubes (SWNTs). The finite temperature effect is accounted for via the local harmonic approach. The equilibrium SWNT configurations are obtained by minimizing the Helmholtz free energy with respect to seven primary coordinate variables that are subjected to a chirality constraint. The calculated specific heats agree well with the experimental data, and at low temperature depend on the tube radii with small tubes having much lower values. Our calculated coefficients of thermal expansion (CTEs) are universally positive for all the radial, axial and circumferential directions, and increase with increasing temperature. The armchair tubes see very large circumferential CTEs, while the zigzag tubes see very large axial CTEs. The tube chirality affects mostly the axial and the circumferential CTEs, but not the radial CTEs. 1 Introduction Carbon nanotubes (CNTs) possess high stiffness and strength and low aspect ratio and density. These extraordinary mechanical properties arouse tremendous interests in CNTs-based nanocomposites [Srivastava & Atluri (2002), Chung & Namburu (2004), Shen & Atluri (2004), Nasdala, Ernst & Lengnick (2005), Gao & Gao (2005)]. Thermal conductance and expansion of the CNTs are two key properties influencing the mechanical behaviors of the nanocomposites in manufacturing and operation. Electrically, a single wall carbon nanotube (SWNT) can be either metallic or semi-conducting depending on its chirality, leading to the possibility to create CNT-based nanoscale electronic device components [Maiti (2002)]. The observation that conductance of a metallic CNT changes by orders in magnitude when strained also opens the door to the potential application of strain-tuned nanoscale electronic transducer, transistor and switcher [Yang, Han & Anantram (2002)]. The thermal properties of CNTs also play critical roles in controlling the performance and stability of these nanoscale electronic components [Liew, Wong & He (2005)]. Many research efforts have been made to determine the specific heat of CNTs, both theoretically and experimentally. Yi, Lu, Zhang, Pan & Xie (1999) experimentally indicated that over the temperature range of 10 − 300o K, the specific heat of multiwall carbon nanotubes (MWNTs) follows a linear temperature dependence, which they attributed to the constant phonon spectrum. Their results indicated that the out-of-plane acoustic mode (as in a graphene sheet) dominated the heat capacity. Mizel, Benedict & Cohen (1999) measured the specific heat for MWNTs in the temperature range 1 < T < 200o K and found a quadratic temperature dependence of the specific heat at low temperature (< 50o K) and a linear temperature dependence above that. Hone, Batlogg, Benes & Johnson (2000) and Popov (2002) made similar observations as Mizel, Benedict & Cohen (1999). Cao, Yan & Xiao (2003) calculated the specific heat using a two-atom unit cell model and the lattice dynamics. The specific heat was found to be proportional to the tubule diameter at low temperatures and inversely proportional to the square of the diameter at high temperatures. Zhang, Xia & Zhao (2003) used a continuum based model to calculate the phonon dispersion relations for SWNTs, based on which they found that the axial lattice wave propagations contributed the most to the specific heat. Li & Chou (2005) calculated the specific heat of SWNTs using molecular structural mechanics and showed that the specific heat increased with increasing tube diameter within the temperature range of 25 − 350o K. Studies on the thermal expansion of CNTs are very limited. Due to the difficulty in nanoscale experiments, most of the experiments focused on CNT bundles and ropes. Ruoff & Lorents (1995) suggested that the radial coefficient of thermal expansion (CTE) of MWNTs be 1 Center for Aerospace Research and Education, University of Caliessentially identical to the axial CTE. The radial CTE fornia at Irvine, 5251 California Ave., Suite 140, Irvine, CA 92612 92 c 2006 Tech Science Press Copyright of MWNTs was found to increase with temperature, nearly identical to that of the c-axis thermal expansion of graphite [Bandow (1997)]. The average tube diameter was observed to increase with increasing growth temperature [Bandow & Asaka (1998)]. Maniwa, Fujiwara, Kira & Tou (2000) reported a radial CTE range of 1.6 × 10−5 − 2.6 × 10−5 /K for the MWNTs. The X-ray studies by Yosida (2000) and Maniwa, Fujiwara & Kira (2001) on SWNT bundles suggested negative radial CTE at low temperatures and positive radial CTE at high temperatures. Although the CTE of SWNT is of fundamental importance to both the nanoelectronics and nanocomposites, experimental data are not available on the CTE for individual SWNT. Theoretical investigations of the thermal expansion of SWNTs are also lacking, and sometimes with contradicting results. Raravikar, Keblinski & Rao (2002) performed MD simulations on (5, 5) and (10, 10) nanotubes and reported temperature independent positive values for both the radial and axial CTEs. The MD simulations by Kwon, Berber & Tománek (2004) indicated negative CTES for SWNTs up to 900o K. Jiang & Liu (2004) showed that both the radial and axial CTEs of SWNTs are negative at low temperature but positive at high temperature, but they did not consider the multibody interactions in deriving the atom vibrating frequencies. Lately, the molecular structural approach by Li & Chou (2005) indicated that both the axial and radial CTEs were positive and increase with increasing temperature. In this paper, we endeavor to analyze the specific heats and the thermal expansion based on a lattice-based cell model. In Section 2, the framework of the cell model for calculating the specific heat and the thermal expansion coefficient is presented. In Section 3, we present results and analysis of the calculated specific heats and CTEs versus temperature and tube radii for different tube chiralities. Section 4 summarizes the work. CMES, vol.14, no.2, pp.91-100, 2006 xA = r, yA = zA = 0, where r is the radius of the tube. The polar coordinates of atom B are given by (r, ϕB, zB ), where ϕB = cos−1 xB /r. The positions of atoms C, D are similarly given. The second nearest neighbor atoms are located using the nearest neighbor atom coordinates. For instance, the equivalence of bond BB1 and CA yields ϕB1 = ϕB + (ϕA − ϕC ) = ϕB − ϕC , (1) zB1 = zB + (zA − zC ) = zB − zC . (2) Similarly, the positions of B2, C1, C2, D1, D2 can be derived. (a) Tubular SWNT 2 Cell model for SWNT using the local harmonic approach (b) Unrolled planar structure The cell model for SWNT is illustrated in Figure 1. In Figure 1, the representative atom A is surrounded by three nearest neighbor atoms B, C and D, forming a lattice cell that can be taken as the basic element of the tube. The second nearest neighbor atoms B1, B2, C1, C2, D1, D2 interact with A through multibody atomistic potentials (e.g., the Tersoff-Brenner potential in below). Now we introduce a polar coordinate system such that Figure 1 : Cell model of SWNT in the tubular and unrolled planar structures. SWNTs can be imagined as a rolled graphene sheet. The unrolled planar graphene sheet, as illustrated in Figure 1(b), can be visualized by cutting the SWNT along its axial direction followed by “unrolling” it without stretching 93 Cell model for calculating thermal capacity & expansion of SWNT to the tangent plane at A. In the planar graphene, we set a 2D Cartesian coordinate system such that xA = 0, yA = 0. Then, the positions of the nearest neighbor atoms in the 2D Cartesian system are given by (rϕi , zi ), where i = B,C, D. The graphene basis vectors a1 and a2 are now given by −→ a1 = BD = [r(ϕD − ϕB ), zD − zB ], (3) vibration coupling among different atoms, thus providing an computationally efficient and conceptually simple method to calculate the free energy of a system. In order to calculate the total potential Utot , the interatomic potential is introduced herein. For carbon atoms, we employ the Tersoff-Brenner potential [Brenner (1990)], which is expressed as Vi j = VR(ri j ) − B i jVA (ri j ), (9) −→ a2 = CD = [r(ϕD − ϕC ), zD − zC ] . (4) for atoms i and j, where ri j is the distance between them. The repulsive and attractive terms are given by In carbon nanotubes (CNTs), the graphene is rolled up in such a way that a graphene lattice vector c = na1 + ma2 D(e) −√2Sβ(r−R(e) ) fc(r), (10) becomes the circumstance of the tube, where the chirality VR (r) = S − 1 e (n, m) uniquely determines the tube. Utilizing the equations (3) and (4), the circumstance of the tube is now D(e) S −√2/Sβ(r−R(e)) given by e fc (r), (11) VA (r) = S−1 √ |c| = c · c where the function fc is a smooth function used to limit r2 [n(ϕD − ϕB ) + m(ϕD − ϕC )]2 = the range of the potential to neighbor atoms, the nearest 1 (1) i.e., f c (r) = 1, 2 {1 + cos π(r − R )/(R(2) − R(1)) }, 0 2 + [n(zD − zB ) + m(zD − zC ) ] . (5) for r < R(1), R(1) < r < R(2), r > R(2) , respectively. The parameter B i j takes account of the multibody interaction Meanwhile, through the bond angles formed at atom i, and is given |c| = 2πr. (6) by 1 2 Equations (5) and (6) yield g = r2 [n(ϕD − ϕB ) + m(ϕD − ϕC )]2 − 4π2 + [n(zD − zB ) + m(zD − zC ) ]2 ≡ 0, B i j = (Bi j + B ji ) , (7) where (12) where g is a geometric constraint that connects the tube Bi j = 1 + ∑ G θi jk fc (rik k(=i, j) chirality to the coordinate variables. LeSar, Najafabadi & Srolovitz (1989) proposed the local harmonic (LH) approach to calculate the Helmholtz free energy for a finite-temperature equilibrium atomic solid. At the heart of the LH approach is the local description of the atomic vibrations, i.e., 2 2 1 ∂ U tot = 0, i = 1, 2, . . . N, ω I3×3 − (8) iκ mC ∂xi ∂xi where mC is the carbon atom mass, Utot is the total potential energy of the system, and ωi is the vibrating frequencies of atom i (varying from 1 to N – the total number of atoms) in the κ(= 1, 2, 3) direction determined with the rest atoms fixed at their equilibrium positions (as implied by the partial derivatives). The LH model neglects the −δ , c20 c2 , G (θ) = a0 1 + 02 − 2 d0 d0 + (1 + cos θ)2 (13) (14) 2 − r2jk /2ri j rik defines the anand where cos θi jk = ri2j + rik gle subtended by the adjoint carbon bonds i − j and i − k. The material parameters employed in this paper are given in the Appendix. The Helmholtz free energy using the local harmonic model is now given as [LeSar, Najafabadi & Srolovitz (1989), Foiles (1994)]: N 3 hωiκ , (15) H = Utot + kB T ∑ ∑ ln 2 sinh 4kB T i=1 κ=1 94 c 2006 Tech Science Press Copyright CMES, vol.14, no.2, pp.91-100, 2006 where kB and h are the Boltzmann and the Planck’s constants, respectively. The total potential Utot directly influenced by a change of atom A’s position is reflected in those bonds of the first closest layers, i.e., AB, AC, AD, and through the bond angles in those of the second closet layers, i.e., BB1, BB1,CC1,CC2, DD1, DD2. Therefore, the total energy can be expressed as Utot = VAB +VAC +VAD +VBB1 +VBB2 +VCC1 +VCC2 +VDD1 +VDD2 where 1 Ua = (VAB +VAC +VAD ) (20) 2 is the potential energy per atom. Therefore, the Helmholtz free energy per atom Ha for the LH approach can be expressed as 3 hωAκ Ha = Ua + kB T ∑ ln 2 sinh . (21) 4kb T κ=1 The equilibrium atom positions for SWNTs at homoge(16) nous finite temperature T can be obtained by solving for the minimum of Ha , i.e.: Jiang & Liu (2004) erroneously neglected the multibody energy contributions represented by the second line on ∂Ha = 0, ∂Ha = 0, and ∂Ha = 0, (22) ∂zi ∂r the right-hand side of equation (16), as the second nearest ∂ϕi neighbor atoms (B1, B2, C1, C2, D1, D2) could not be where i = B,C, D. Note that the minimization is subcharacterized in their model. jected to the nonlinear chirality constraint g ≡ 0. The vibrating frequencies at atom A can now be derived Once the equilibrium configuration is solved, the specific by substituting the total energy Utot (16) into equation heat per mass is given by [Jiang & Huang (2005)] (8). In doing so, we point out the seven primary variables d ln ωAκ 1 3 − ω2Aκ T dT kB (namely, unknowns) in our model, i.e., (ϕB , zB ), (ϕC , zC ), (23) Cv = ∑ sinh2 (ωAκ ) , (ϕD , zD) and r. The second nearest atoms are specified mC κ=1 using the bond equivalence as represented by equations where (1) and (2). It can be readily proved for a SWNT of a homogenous temperature T , the frequencies ωiκ are in- ωAκ = hωAκ , (24) 4πkB T dependent of the atom i and can be taken the same as those of the representative atom A. However, we need is the dimensionless frequency. For numerical conveto mention that although the global diagonalization from nience, (23) is approximated using a backward difference the quasiharmonic to the local harmonic approaches de- scheme. couples the vibrating among atoms, the directional cou- The thermal expansion is characterized by the coefficient pling within the local harmonic model cannot be readily of thermal expansion, defined as taken as null. Hence, the off-diagonal component in the 1 dli , (25) local dynamic matrix cannot be neglected as did in Jiang αi = li dT & Huang (2005). where li is the instantaneous length given by Now the Helmholtz free energy of the system at finite temperature can be obtained as lz = max (zB − zC , zD − zC ), 3 hωAκ , (17) lr = r, H = Utot + kB T N ∑ ln 2 sinh 4kB T κ=1 + bond energies independent of atom A. where ωAκ is also a function of the primary variables. The total potential energy Utot for the Tersoff-Brenner formalism can be written as 1 (18) Utot = ∑ ∑ Vi j . 2 i j=i Using the bond equivalence, it can be shown that Utot = NUa , (19) lc = r max (ϕD − ϕB , ϕD − ϕC ) , where z, r, c represent respectively the axial, radial and circumferential directions. In our numerical implementations, a central finite difference scheme is employed to approximate (25), i.e., αi = 1 liT +ΔT − liT −ΔT . 2ΔT liT (26) 95 Cell model for calculating thermal capacity & expansion of SWNT 3 Results and discussions 1500 Cv (mJ/g − K) Figure 2 shows the calculated Cv for the armchair (n, n) SWNTs. In Figure 2(a), for comparison purpose, the specific heats for graphite and diamond are also shown. In Figure 2(b), the calculated Cv ’s are shown for the low temperature range of 2 − 300o K, together with Hone et al. [Hone, Batlogg, Benes, et al. (2000)] measured data for SWNT ropes. Over a wide range of temperature simulated, especially for high temperature above 100o K, the present analysis agrees well with the experimental data. The results indicate that the specific heats do not depend on the radius of the nanotubes, except at temperature range of 2 − 300oK. This high temperature independence of the specific heat was attributed to the phonon states of the constituent graphene sheet [Hone, Laguno & Biercuk (2002)]. The calculated high temperature specific heats approach the theoretical limit value of 2078mJ/g − K, regardless of the the chirality and radius of the tube. 1000 Charility (5,5) (10, 10) (20, 20) 500 Exp. data for graphite [32] Exp. data for diamond [33] [34] Exp. data for diamond 0 0 500 T (o K) 1000 (a) 2 − 1000K Cv (mJ/g − K) 600 Figure 3 shows the specific heats versus the radius of Charility (5,5) the nanotubes at three levels of temperature. The open (10, 10) 500 symbols represent the data for the armchair tubes, while (20, 20) the the filled tubes represent those for the zigzag tubes (25, 25) 400 (5, 0), (10, 0), (20, 0), (30, 0). Figure 3 further conExp. data [12] firms that the specific heats are radius independent at 300 high temperature, and that at low temperature, the specific heats show a strong dependence on the tube ra200 dius for small tubes and a weak dependence on the tube radius for large tubes. No obvious differences are ob100 served for the armchair and zigzag tubes. We also calculated the Cv for different tube chiralities (25, m), where 0 m = 0, 3, 6, 9, 15, 20,25. Our results (not shown here) in0 100 200 300 dicate that Cv is only very slightly dependent on the chiT (o K) ralities of the tubes and that Cv ’s for the chirality tubes (b) 2 − 300K (with m = 3, . . ., 20) are contained in those of the armchair and the zigzag tubes. Figure 2 : Temperature dependence of specific heat for Our results deviates below the experimental data for tem- (n, n) SWNTs. perature lower than 100oK. Figure 4 shows the contributions to the specific heat from the three atom vibrating modes for (10, 10) tube. As observed by Yi, Lu & Zhang (1999), at low temperature, our results clearly show that (expect κ = r) is quite independent on the radius of the the out-of-plane vibrating mode (the radial mode) domi- tubes. In Figure 5, we also plot the horizontal line of nates the specific heat of the SWNT. As the temperature ln(2 sinhω) = 0, namely, ω = 0.48. Above about 500o K, increases, the contributions from the circumferential and ωr turns to increase, instead of decreasing, the Helmholtz then the axial vibrating modes gradually increase. Fig- free energy. The radial mode is unique to the SWNTs ure 5 gives the normalized atom vibrating frequencies [Ravavikar, Keblinski & Rao (2002)]. The discrepancy ωκ for the armchair tubes (n, n). It can be seen that ωκ of the calculated Cv at low temperature is caused by the c 2006 Tech Science Press Copyright 96 CMES, vol.14, no.2, pp.91-100, 2006 2000 1400 1200 1000 Total 800 Axial Circumferential 1500 Cv (mJ/g − K) Cv (mJ/g − K) 600 T 400 100 300 200 500 Radial 1000 500 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 1.8 0 200 400 600 800 1000 1200 1400 T (o K) r(nm) Figure 3 : Radius dependence of specific heats for Figure 4 : Mode contributions to specific heats for (10, 10) tube. SWNTs. overestimated radial vibrating frequency, which should be much more constrained by the compressed out-ofplane π bonding orbitals. In the original formalism of Brenner’s potential [Brenner (2000)], the π bond is reflected in the multibody term B i j as Chirality (5,5) (10,10) (20,20) 101 (27) where DH Biπj = ΠRC i j + Bi j , 2 ωκ 1 2 B i j = (Bi j + B ji ) + Biπj , 10 ωz ωc (28) and where the first term ΠRC i j represents the influence of radical energetics and π bond conjugation on the bond energies, and the second term BiDH j depends on the dihedral angle for carbon-carbon double bonds. We tried a constant Biπj = −0.0243 (taken from Brenner [Brenner (1990)] for grahpite) in our calculations. The results overpredicted the low temperature specific heat (comparing to the experimental data) by nearly two times, but approached the experimental values at high temperature (starting at ∼ 300o K). Due to lack of Biπj data and their derivatives for SWNTs, we made no further attempt in enhancing the calculated low temperature specific heat. Figure 6 shows the calculated CTE for the armchair, zigzag and chirality tubes. In each case, the axial and circumferential CTEs are also zeroes at 2o K, but being shifted an amount upwards for clarity. In accordance 10 10 0 ωr -1 0 500 1000 1500 T (o K) Figure 5 : Normalized atom vibrating frequencies for the armchair tubes. with Li et al.’s [Li & Chou (2005)] calculations, our results indicate universal positive CTEs for the radial, axial and circumferential directions. As can be seen from 6(a), for the armchair (n, n) tubes, the smaller tubes have slightly higher axial but lower radial CTEs than the larger tubes. The two large tubes (20, 20) and (25, 25) shows 97 Cell model for calculating thermal capacity & expansion of SWNT 1E-05 CTE 8E-06 αc 6E-06 4E-06 Chirality αz (5,5) (10,10) 2E-06 (20,20) αr 0 0 200 400 (25, 25) 600 800 T (o K) 1000 1200 1400 1000 1200 1400 (a) (n, n) Chirality (5,0) 1.6E-05 (10,0) (20,0) (30,0) 1.2E-05 CTE The CTE variations for the zigzag tubes are shown in Figure 6(b). For the zigzag tubes, the tube radius has a more pronounced effect on the thermal expansion. The large tube (30, 0) has a radial CTE that is nearly two to three times larger than that of the small tube (5, 0). But the axial CTE for (30, 0) is about two to three times lower than that for the small tube (5, 0). Very high axial CTEs are observed for the small zigzag tube. Note that the axial direction for the zigzag tubes is the circumferential direction for the armchair tubes. Hence, this observation of high axial CTEs for the zigzag tubes is in accordance with the above observation of high circumferential CTEs for the armchair tube. The explanation for this preference of thermal expansion is simple. Take the armchair tube (10, 10) for instance. Table 1 gives the atom positions in the 2D Cartesian system for T = 300, 1000oK. It is seen that atom B moves slightly away from atom A (at xA = yA = 0) nearly along the circumference, and that C moves slightly away from A nearly along the axial direction. But atom D is also seen to move slightly away along the circumference. Bond AD’s circumferential extension, together with bond AB circumferential extension and counterclockwise rotations, causes the large circumferential CTE for the armchair tubes. For the zigzag tubes, no paradox is seen for the circumferential CTEs, which now follows nearly identical paths as the radial CTEs. 1.2E-05 8E-06 αz 4E-06 0 αc αr 0 200 400 600 800 T (o K) (b) (n, 0) 1E-05 8E-06 CTE almost the same CTEs, which means that the CTEs are radius independent for large tubes. To the best the author’s knowledge, the circumferential CTE is for the first time reported in the literature. Interestingly, for the armchair tubes, the radial and the circumferential CTEs do not follow the same path. For instance, the small tube (5, 5) has a lower αr but a much higher αc than those of (10, 10), although one might expect αc = αr . This seemingly contradiction reflects the self-adjusting of the unit cell composed of atoms A, B, C, D in the minimization process of the Helmholtz free energy. For the small tubes, the paradox suggests more self-extension in the circumference. 6E-06 αz 4E-06 αc Chirality (25,0) 2E-06 (25,9) αr 0 0 (25,25) 200 400 600 800 1000 1200 1400 Figure 6(c) shows the effect of the chirality on the therT (o K) mal expansion. Opposite temperature effects are seen for (c) (25, m) the axial and the radial CTEs. Comparing the armchair and the zigzag tubes, the zigzag tube (25, 0) has a much Figure 6 : Temperature and chirality dependence of larger axial CTE and a relatively lower radial CTE. And CTEs for SWNTs. as seen above, the armchair tube has a high circumferential CTE. We need to mention that small local oscillations 98 c 2006 Tech Science Press Copyright CMES, vol.14, no.2, pp.91-100, 2006 Table 1 : Equilibrium atom positions for (10, 10). The length unit is 10−2nm . 300 1000 B (-7.294, 12.597) (-7.330, 12.600) C (-7.267, -12.635) (-7.264, -12.688) D (14.603, 0) (14.640, 0) are seen on the calculated CTEs. This small local oscilla- Appendix tions are introduced by the built-in convergence criteria The parameters for Brenner’s potentials [Brenner of the IMSL optimizaiton package we used. Small and (1990)]: artificial, they are smoothed out by a polynomial fitting. D(e) = 6.0eV, R(1) = 0.17nm, S = 1.22, β = 21nm−1 , R(2) = 0.27nm, a0 = 0.00020813, c0 = 330, R(e) = 0.139nm, d0 = 3.5. 4 Summary In summary, we develop a lattice-based cell model to calculate the specific heats and the coefficients of thermal expansion of single wall carbon nanotubes. The cell model consists of seven primary coordinate variables while subjecting to a chirality constraint. The specific heat and thermal expansion of the SWNTs are studied. The calculated specific heats are in good agreement with experimental data and for a large range of temperatures, very close to those of graphite and diamond. The chirality dependence of the specific heat is seen only up to a few hundred Kelvins. Small tubes have much lower values at low temperature. Positive thermal expansions are observed for all the radial, axial and circumferential directions. The coefficients of thermal expansion (CTEs) increase with increasing temperatures. For the armchair tubes, both the radial and axial CTEs are only slightly influenced by the tube radii. But the armchair tubes are seen to have large values for the circumferential CTEs. The zigzag tubes have small radial and circumferential CTEs, which also weakly depend on the tube radii. But the zigzag tubes see very large axial CTEs, which also strongly depend on the tube radii. The axial and circumferential CTEs, but not the radial ones, are much influenced by the tube chiralities. Acknowledgement: This work is supported by the Army Research Laboratory. The first author would also like to thank Dr. Brenner for providing his nanotube code that generates the zero temperature (n, m) carbon nanotubes. The carbon bond length for the zero temperature graphene sheet is set to 0.14507nm. The other constants are given as: mc = 1.9926 × 10−26Kg. The Planck’s constant h = 6.626068 × 10−34Js. And the Boltzmann constant kB = 1.3806503 × 10−23JK −1 . References Allen, M.; Tildesley, D. (1987): Computer Simulation of Liquids. Oxford University Press. Srivastava D., Atluri S.N. (2002): Computational Nanotechnology: A current perspective, CMES: Computer Modeling in Engineering & Science, vol. 3, pp. 531-538. Chung P.W., Namburu R.R., Henz B.J. (2004): A Lattice Statics-Based Tangent-Stiffness Finite Element Method, CMES: Computer Modeling in Engineering & Science, Vol. 5, pp. 45-62. Shen S.P., Atluri S.N. (2004): Computational Nanomechanics and Multi-scale Simulation, CMC: Computers, Materials & Continua, vol. 1, pp. 59-90. Nasdala L., Ernst G., Lengnick M. & Rothert H. (2005): Finite Element Analysis of Carbon Nanotubes with Stone-Wales Defects, CMES: Computer Modeling in Engineering & Science, Vol.7, pp. 293-304. Guo W.-L. and Gao H.-J. (2005): Optimized Bearing and Interlayer Friction in Multiwalled Carbon Nanotubes, CMES: Computer Modeling in Engineering & Science, Vol. 7, pp. 19-34 Cell model for calculating thermal capacity & expansion of SWNT 99 Maiti A. (2002): Select applications of carbon nanotubes: field-emission devices and electromechanical sensors, CMES: Computer Modeling in Engineering & Science, vol. 3, pp. 589-599. Bandow S., & Asaka S. (1998): Effect of the growth temperature on the diameter distribution and chirality of Single-Wall Carbon Nanotubes, Phys. Rev. Letters, vol. 80, pp. 3779-3782. Yang L., Han., Anantram M.P. (2002): Bonding geometry and bandgap changes of carbon nanotubes under uniaxial and torsional strain, CMES: Computer Modeling in Engineering & Science, vol. 3, pp. 675-685. Maniwa Y., Fujiwara R., Kira H., & Tou H. (2000): Multiwalled carbon nanotubes grown in hydrogen atmosphere: An x-ray diffraction study, Phys. Rev. B, vol. 64, 073105. Liew K.M., Wong, C.H., He X.Q., & Tan T.J. (2005): Yosida Y. (2000): High-temperature shrinkage of singleThermal stability of single and multi-walled carbon nan- walled carbon nanotube bundles up to 1600 K, Journal of Applied Physics, vol. 87, pp. 3338-3341. otubes, Phys. Rev. B, vol. 71, 075424. Yi W., Lu L., Zhang D.L., Pan Z.W., & Xie S.S (1999): Maniwa Y., Fujiwara R., Kira H., & Tou H. (2001): Linear specific heat of carbon nanotubes, Phys. Rev. B, Thermal expansion of single-walled carbon nanotube (SWNT) bundles: X-ray diffraction studies, Phys. Rev. vol. 59, 9015. B, vol. 64, 241402 (R). Mizel A., Benedict L.X., Cohen M.L., et al. (1999): Analysis of the low-temperature specific heat of mul- Raravikar N.R., Keblinski P., & Rao A.M. (2002): tiwalled carbon nanotubes and carbon nanotube ropes, Temperature dependence of radial breathing mode Raman frequency of single-walled carbon nanotubes , Phys. Phys. Rev. B, vol. 60, 3264?270. Rev. B, vol. 66, 235424. Hone J., Batlogg B., Benes Z., Johnson A.T., Fischer Kwon Y.-K., Berber S., & Tománek (2004): Thermal J.E. (2000): Quantized phonon spectrum of single-wall Contraction of Carbon Fullerenes and Nanotubes, Phys. carbon nanotubes, Science, vol. 289, pp. 1730-1733. Rev. Letters, vol. 92, 015901. Popov V.N. (2002): Low-temperature specific heat of Jiang H., Liu B., et al. (2004): Thermal expansion of nanotube systems, Phys. Rev. B, vol. 66, 153408. single wall carbon nanotubes, J. Engng. Mater. Tech., Cao J.X., Yan X.H., Xiao Y. (2003): Specific heat v126, pp. 265-270. of single-walled carbon nanotubes: a lattice dynamics Li C.Y., Chou T.-W. (2005): Axial and radial thermal study, J. Physical Society of Japan, vol. 72, 2256-2259. expansion of single-walled carbon nanotubes , Phys. Rev. B, vol. 71, 235414. Zhang S.L., Xia M.G. & Zhao S.M. (2003): Specific heat of single-walled carbon nanotubes, Phys. Rev. B, LeSar R., Najafabadi R., Srolovitz D.J. (1989): Finitevol. 68, 075415. temperature defect properties from free-energy minimization, Phys. Rev. Letters, vol. 63, pp. 624-627. Li C.Y., Chou T.W. (2005): Quantized molecular structural mechanics modeling for studying the specific heat Brenner D.W. (1990): Empirical potential for hydrocarof single-walled carbon nanotubes, Phys. Rev. B, vol. 71, bons for use in simulating the chemical vapor deposition 075409. of diamond films, Phys. Rev. B, vol. 42, pp. 9458?471. Ruoff R.S., & Lorents D.C. (1995): Mechanical and Foiles S.M. (1994): Evaluaiton of harmonic methods for thermal properties of carbon nanotubes, Carbon, vol. 33, calculating the free energy of defects in solids, Phys. Rev. pp. 925?30. B, vol. 49, 14930. Bandow S. (1997): Radial thermal expansion of purified Jiang H., et al. (2005): A finite temperature continmultiwall carbon nanotubes Measured by X-ray diffrac- uum theory based on interatomic potentials, Trans. of the ASME, vol. 127, pp. 408-416. tion, Jpn. J. Appl. Phys., vol 36, pp. L1403-L1405. 100 c 2006 Tech Science Press Copyright Hone J., Laguno M.C., Biercuk M.J., et al. (2002): Thermal properties of carbon nanotubes and nanotubebased materials, Appl. Phys. A, vol. 74, pp. 339-343. Madelung L.-B., et al. (1982): Physics of Group IV elements and III-V compounds, New series, Group III, vol. 17, Springer-Verlad, Berlin. Billings B.H., & Gray D.E. (1972): American Institute of Physics Handbook, McGraw-Hill, New York. Pitzer K.S. (1938): The heat capacity of diamond from 70 to 300o K, Journal of Chemical Physics, vol. 6, pp. 68-70. Ravavikar N.R., Keblinski P., Rao A.M. (2002): Temperature dependence of radial breathing mode Raman frequency of single-walled carbon nanotubes, Phys. Rev. B, vol. 66, 235424. Brenner D.W. (2000): The art and science of an analytic potential, Phys. Stat. Sol. (b), vol. 217, pp. 23-40. CMES, vol.14, no.2, pp.91-100, 2006